DOI:
10.1039/B003188G
(Paper)
J. Mater. Chem., 2001,
11, 149-152
Transformation of porous doped silica xerogels
into multicomponent bulk amorphous monoliths with interesting optical properties†
Received 19th May 2000, Accepted 18th July 2000
First published on 19th October 2000
Abstract
We have developed a technique for the fabrication of glassy silica
materials with ultrafine metal and semiconductor particles. The porosity of
xerogels produced under incomplete polycondensation strongly influences the
chemical state of the final products, which are formed by reactions of xerogel-dopants
in different atmospheres. We have studied, in particular, the process leading
to copper sulfide and copper selenide nanoparticles which are promising for
non-linear optical applications due to the appearance of a new absorption
band in the near IR.
1.0 Introduction
Porous solids have a number of unique properties due to their large active
interface with the atmosphere and the peculiar structure of pores usually
formed spontaneously1,2 or by means of
some lithographic method.3 They can be both
inert supports (e.g., containing a catalyst for reactions with gaseous
or liquid components) and play an active role, providing chemical processes
with participation of species bound to the material (e.g., ion-exchange
in zeolites). Silica is one of a number of porous solids and has been extensively
studied for different purposes, however, there are still unresolved questions
even in this quasi-monocomponent material. An ideal chemical structure
with tetrahedral SiO4 groups is not an adequate representation
of any porous silica because of its amorphous character and numerous surface
hydroxyl groups arising either from synthesis or chemisorbed from the atmosphere.
Meanwhile, the porosity allows the combination of silica with a wide variety
of other components which may be introduced into the pores with further conversion
into a desired dense state. One such procedure which combines porous silica
and transition metal ions was designed by us for the fabrication of optical
materials with unique optical properties. Fortunately, silica is very suitable
for applications in optics due to its good transparency through the UV/Vis/near
IR range, and its chemical reactivity is low enough to avoid any chemical
interaction of the matrix with dopants. The familiar sol–gel process4–6 includes a porous silica xerogel
step which provides a useful method for silica doping and modification. Conditions
of further transformations allow control of the chemical state of such an
X-SiO2 material. As X in this paper we considered copper, copper
oxides and chalcogenides; however, many other elements and compounds (each
requiring suitable conditions) could be analogously introduced into the porous
silica xerogel. One other exciting feature of this process is the production
of dopants in the form of small particles dispersed within the silica matrix:
the particles can be metallic or semiconducting comprising “nanoparticles-in-dielectrics”
materials with quantum confinement effects in the nanoparticle properties
and particle–particle interactions.7–9The purpose of the present work was the fabrication of multicomponent silica-based
solid materials through porous xerogels with ultrafine copper compounds, and
study of their structural and optical properties. The materials fabricated
are of interest, in particular, for optical element applications, and the
method presented allows variation of their characteristic features for the
construction of filters, non-linear switches, and optical limiters. Fabrication
routes and some physical properties of sol–gel derived silica materials
with copper and copper oxides were reported previously for monolithic compositions10–13 and films;14,15
however, our research was aimed at materials with semiconductor copper chalcogenide
particles produced in a glassy optically transparent silica matrix rather
than every CuX–SiO2 nanocomposite.
2.0 Preparation and further processing
The complete cycle of the sol–gel procedure of multicomponent matrix
formation (from liquid sol to stiff transparent glass) partially repeats the
familiar routes16–21
which were suggested for optical materials doped with small semiconductor
particles; however, fabrication of the sol–gel glasses with copper chalcogenides
was performed by us for the first time and published in ref.
22. The preparation route includes the following sequence of steps.
The precursor solution was prepared by mixing of tetraethoxysilane (TEOS),
water, and HCl (mole ratio 1 ∶ 25 ∶ 0.05,
respectively). In order to overcome a strong volume contraction during gel
drying, aerosil with a particle size of 15–20 nm was added to
the sols. It should be noted that approx. 50% contraction remained with the
aerosil present, but in its absence no stiff and crack-free xerogels could
be obtained with the given procedure. After mixing and stirring for 1 h,
ultrasonic activation and removal of agglomerates from the sol by centrifugation,
aqueous ammonia was added to increase the pH of the sol to 7–8. Then
this sol was poured into closed polystyrene containers and left for a day
for gelation. Gels were heated at 800
°C for 1 h and uniform
porous samples in the form of disks were produced. These xerogels were used
as initial starting materials for the preparation of samples doped with different
copper compounds. Copper was introduced into the xerogels by impregnation
in alcoholic solutions of Cu(NO3)2 for 8 h. Impregnated
doped xerogels were dried in air and heated at 600
°C for 1 h.
The next operations with the copper-doped xerogels included one of the
routes described below, and shown schematically in Fig. 1. |
| Fig. 1 The preparation sequence
of silica monoliths doped with metallic copper, copper sulfide, copper oxide
and copper selenide. A detailed description of each step is given in the text. | |
(i) Annealing in air at a maximum temperature of 1200
°C (for
10 min) led to the formation of a transparent light-brown glass.
The subsequent heating of the sol–gel glass was carried out in flowing
hydrogen at 600
°C for 1 h. As a result of such treatment
the light-brown glasses were transformed into red transparent glasses.
Analogous materials were obtained in one step after the high-temperature
annealing of the copper-doped xerogels in flowing hydrogen (Fig. 1).
(ii) Heating in flowing hydrogen (600
°C for 1 h) followed
by annealing in closed quartz ampoules (ampoule volume 10–20 times greater
than sample volume) together with a small amount of elemental selenium. The
amount of selenium was calculated to provide a partial pressure of Se vapour
of about 1 atm at the maximum temperature of the heat treatment (1200
°C).
The ampoules were sealed without any outgassing of air. The xerogels in the
closed ampoules were heated to 600
°C, held for 1 h at this
temperature and further heated to 1200
°C with a temperature ramp
rate of 100
°C h−1, and finally heated
at 1200
°C for 10 min. As a result of this sequence the
xerogels resulting from treatment (i) were also transformed into transparent
dark-green glassy samples.
(iii) Heating in flowing hydrogen sulfide (400
°C for 1 h)
followed by annealing in sealed quartz ampoules similarly to treatment (ii),
but without selenium. The ampoules were also sealed without any outgassing
and heated by the same temperature regime to 1200
°C. This route
resulted in transparent dark-brown glassy samples. All glasses fabricated
were polished to an appropriate thickness for optical measurements. Thus,
we studied four types of glassy samples (monoliths) transformed from Cu-doped
xerogels: oxidized, hydrogen reduced, selenized and sulfidized. We expected
the appearance of copper oxide, metallic copper, copper selenide and copper
sulfide, respectively, within these materials after the above described procedures.
3.0 Results and discussion
3.1. X-Ray diffraction
The glassy materials under study were produced via the transformation
of a porous xerogel into a transparent glassy monolith. This transformation
is inherent to the pure xerogels which were not doped with copper or with
any other component. A dopant is required to be inert at the loading used
(the admissible maximum amount has been established experimentally and in
the typical samples presented was 0.5–1 at.%). Note that the
annealing process can be changed completely when higher amounts of dopant
are used. We kept a low dopant amount, and the heating and annealing regimes
were same for the doped and pure xerogels. These regimes resulted in the formation
of high optical quality, transparent samples in the final steps. This allowed
us to study the annealing process with pure xerogels. In other words, it was
assumed that the behaviour of the silica matrix of the pure xerogels is same
as in the doped ones. X-Ray diffraction (XRD) analysis was carried out
for a series of xerogels heated under different temperatures (600–1250
°C)
(Fig. 2). The first set of XRD data
(Fig. 2a) characterizes xerogels during
the pre-annealing steps: they remain very porous and no volume contraction
occurred. The second set (Fig. 2b) corresponds
to the annealing observable by the naked eye: a volume contraction (up to
50%) with conservation of the initial shape (i.e. the samples were
not melted) and the appearance of transparency, with an optical spectrum similar
to fused silica glass formed by traditional methods. These diffractograms
exhibit the well-known broad peak of amorphous silica, the position of
which is practically identical for both sets of xerogels (i.e. those
heated to 600–1000
°C and 1100–1250
°C).
The maximum of this peak appears at 2θ = 23°
with a peak width from 15 through 30°; it originates from the disordered
–O–Si–O– network of the amorphous material and may
be related to the contribution from the first and second neighbouring Si atoms.
The degree of this disorder is slightly increased in the xerogels from 600
to 1000
°C, and it is retained under the different steps of annealing.
No traces of crystallization were observed at this temperature. The cristobalite
phase of SiO2 was formed only after heat treatment at 1500
°C
(not shown in Fig. 2). It should be
kept in mind that the radial distribution function of atoms may not be evaluated
from wide-range XRD, however, the data suggest that the local structure
of silica does not change significantly. The annealed xerogels are the same
amorphous material as the porous one. The annealing process only results in
a decrease in the pore volume (until the disappearance of the pores). |
| Fig. 2 A series of X-ray
diffraction data in the range of 2θ = 15–35°
for the pure silica xerogels and monoliths (i.e. annealed xerogels)
without any doping after heating at different temperatures (specified on the
curves) for 1 h. The ordinate is given in arbitrary units. | |
Note that the Cu-doped annealed xerogels have similar diffractograms:
the amount of copper atoms in silica glass is probably too small to be detectable
with XRD, but the local silica structure has not been appreciably changed
by the presence of copper.
3.2. Optical spectroscopy
As the result of the xerogel-to-glass transformation according
to route (iii), coloured glassy samples were obtained which were studied with
optical absorption spectroscopy. Fig. 3
displays typical spectra for the four kinds of materials synthesised. The
almost structureless curve (1) corresponds to the copper oxide step. It is
similar to the spectra of this compound dispersed in films and glasses23,24 and compatible with the indirect band-gap
character of this semiconductor; Eg is estimated to be
1.4–1.5 eV.25,26 A rather
pronounced feature is observed for the hydrogen-reduced monoliths (curve
2, Fig. 3). A peak at 2.1 eV
with a shoulder in the range of 1.7–1.8 eV is typical for ultrafine
copper particles exhibiting a plasmon resonance27
just in this region. Copper particles with a well-expressed plasmon resonance
have been produced within oxide matrices.11,28–30
The exact shape of the resonance depends on the size and sphericity of the
particles and their interactions with the matrix and environment. The latter
may be responsible for the low-energy part of this feature. As a result
of sulfidization and selenization of copper oxide and copper, respectively,
with subsequent annealing, we obtained absorption spectra comprising two parts:
a monotonous absorption rise in the high-energy part with the appearance
of the broad band for selenide samples, and a broad maximum in the low energy
part (curves 3 and 4 in Fig. 3). The
maximum corresponding to sulfide was located at about 1.9 eV, and in
the case of selenide it is shifted to the near-IR region, 1.1 eV.
This rather different position of the maximum could not be associated with
formation of copper ions in the glasses under annealing, absorption of these
covers the range 1.5–2.0 eV,11,31,32
since we observe the explicit effect of chalcogen type species. The interpretation
of this effect as arising from the absorption of ultrafine copper selenide
and sulfide particles is more likely. The fundamental absorption edges of
sulfides with different stoichiometries were determined to be from 1.2 eV
for Cu2S33,34 to 2 eV
for CuS.35,36 For copper selenides the
data on Eg vary from 1.4 to 2.2 eV, depending on
the preparation details and the stoichiometry.37,38
For copper sulfides the properties of quantum-size particles with a similar
maximum were reported in refs.
39–41. These data support the conclusion drawn on the appearance
of copper sulfide particles in our case. The near-IR absorption for selenide
cannot be treated as a quantum-size effect, which would result in a blue-shift
rather than the red-shift observed here. A contribution from quantum-sized
selenide particles can be considered to explain the band between 2.0 and 2.5 eV,
but the near-IR maximum has, perhaps, another origin. Analogously with
the properties of the sol–gel silica films containing copper selenide
(for different stoichiometry, CuxSe, 1 ≤ x ≤ 2)
we propose a partial oxidation of the particles with the formation of interband
levels responsible for transitions with an energy of about 1.1 eV.
A large surface/volume ratio in the case of ultrafine particles provides the
considerably high intensity of this maximum. |
| Fig. 3 Absorption spectra
of the four types of Cu-doped silica monoliths obtained at the steps of
formation of metallic copper (2), and copper compounds: CuO (1), CuxS
(3), and CuxSe (4). | |
3.3. Transmission electron microscopy
From the optical data we can conclude that the different chemical compositions
of the glasses produced by means of chemical transformations of a series of
copper compounds reveal very distinct optical absorption properties. The interpretation
of this fact given above as arising from the occurrence of nanoparticles was
confirmed directly with transmission electron microscopy (TEM). The micrographs
were obtained by means of the ‘replica with extraction’ method
(a carbon film of 10 nm thickness was evaporated onto slightly etched
samples and was separated at the water–air interface by careful immersion
into water and transferred onto standard TEM copper meshs). They indicate
unambiguously that the colour of the glasses is controlled by particles which
are present in rather low concentrations, 1013–1014 cm−3
(the amount of matter is in agreement with the optical absorption intensity).
The particles are near-spherical and their average sizes are collected
in Table 1. These data indicate
that the range of particle sizes is similar in glasses of different compositions,
apart from CuO/SiO2 for which the presence of some amorphous phase
cannot be excluded. Optical spectra for this glass did not reveal any pronounced
features. In the semiconductor-doped, CuxS/SiO2
and CuxSe/SiO2 materials the size of the particles
could contribute only as a weak quantum confinement effect (since typical
values of the Bohr radius of similar medium-band chalcogenide semiconductors
are less than 10 nm42,43). Thus,
the multicomponent materials fabricated can be related to the particles-in-dielectric
matrix systems with explicit separation of the component features.
Table 1 Averaged sizes of nanoparticles formed in
the monoliths according to TEM data
Type of monolith | Particle size/nm |
---|
Under the above TEM observation
using the “replica with extraction” method we did not detected
with any certainty particles less than 5–10 nm; however, other
studies of these samples (kindly performed by Dr. Shixin Wang, Michigan University)
with HRTEM have confirmed the absense of explicit crystalline phases in these
monoliths. |
---|
Cu/SiO2 | 20–50 |
CuO/SiO2 | No particles were observeda |
CuxS/SiO2 | 30–100 |
CuxSe/SiO2 | 20–80 |
3.4. Rutherford backscattering spectroscopy
It should be noted that full analysis of the chemical compositions of similar
composites is very complicated: any destructive methods result in strong distortion
of data since the nanoparticles are not stable in the absence of a glass matrix
or some other protective shell. Our preliminary conclusions on the composition
issue, first of all, from the analogous data on the CuxS–
and CuxSe–silica films, were obtained with XRD and
X-ray photoelectron spectroscopy (XPS) studies.44,45
For the latter, CuxSe/SiO2, the optical properties
of the films and glasses considered in the present work are similar, and we
performed additional compositional studies with Rutherford backscattering
spectroscopy (RBS; Table 2).
Table 2 Data of RBS analysis of CuxSe-containing
silica monoliths within the surface layer of a freshly cleaved sample
Element | Relative amounts (atom%) |
---|
O | 67.82 |
Si | 32.00 |
Cu | 0.13 |
Se | 0.06 |
A typical experimental setup was used: the 1.6 MeV 4He2+
beam incident normally on the samples was detected at a scattering angle of
160°. The data on the elemental composition indicate that the stoichiometry
of the copper selenide formed was Cu2Se, whilst a different selenization
extent can provide a range of selenides CuxSe with controllable x
value within the silica films.46,47
Thus, the analysis performed showed that the properties of these multicomponent
semiconductor–glass materials can be controlled both via the
type of copper compound and the stoichiometry of the chalcogenides. The last
circumstance is inherent to transition metal compounds and is unknown for
the classical semiconductor-doped glasses with AIIBVI,
AIIIBV, and AIVBVI compounds.
The size effect, which is much more usable for control of glass properties,7,8,43,48 was not investigated by
us within the framework of this work, and will be the subject of further studies.
4.0 Conclusions
1. A series of monolithic silica glasses doped with different copper compounds
(oxide and chalcogenides) and metallic copper was fabricated with the sequence
of steps including the sol–gel process, heating and annealing of porous
xerogels in controllable atmosphere for in situ chemical transformations.2. Ultrafine metallic copper, copper sulfide and copper selenide nanoparticles
were formed within the silica matrix at the final annealing step, accompanied
by the transformation of porous xerogels to glasses. The particles, with sizes
in the range of tens of nanometers, which are separately located in the matrix
provide unusual optical properties of the materials.
3. The copper selenide doped glasses are of greatest interest for optical
applications since they possess an intense near-IR absorption band. Their
composition was determined as Cu2Se/SiO2; however, control
of the size of particles and stoichiometry of CuxSe may
be feasible in order to tune the unique optical non-linear properties.
Acknowledgements
The authors thank Drs. A. S. Lyakhov and L. S. Ivashkevich
for assistance in XRD measurements and Dr. K. V. Yumashev and P. V. Prokoshin
for recording of absorption spectra in the near-IR range. This work was
supported by the Belarusian Fundamental Foundation and Ministry of Education
of Belarus.References
- Access in nanoporous
materials, ed. T. J. Pinnavaia and M. F.
Thorpe, Kluwer Acadademic Publishers, Dordrecht,
The Netherlands, 1996. Search PubMed.
- C. N. R. Rao, J. Mater. Chem., 1999, 9, 1 RSC.
- Y. Xia, J. A. Rogers, K. E. Paul and G. M. Whitesides, Chem. Rev., 1999, 99, 1823 CrossRef CAS.
- D. R. Ulrich, J. Non-Cryst. Solids, 1988, 100, 174 Search PubMed.
- L. L. Hench and J. K. West, Chem.
Rev., 1990, 90, 33 CrossRef CAS.
- J.-L. R. Nogues and W.
V. Moreshead, J. Non-Cryst. Solids, 1990, 121, 136 Search PubMed.
- F. Henneberger and J. Puls, in Optics of Semiconductor Nanostructures, Academie
Verlag, Berlin, 1992, p. 497. Search PubMed.
- N. R. Kulish, V. P. Kunets and M. P. Lisitsa, Opt.
Eng., 1995, 34, 1054 CAS.
- L. L. Beecroft and Ch.
K. Ober, Chem. Mater., 1997, 9, 1302 CrossRef CAS.
- M. G. Ferreira Da Silva and J. M. Fernandez Navarro, J. Non-Cryst. Solids, 1988, 100, 447 Search PubMed.
- J. F. Perez-Robles, F. J. Garcia-Rodriguez, J. M. Yanez-Limon, F. J. Espinoza-Beltran, Y. V. Vorobiev and J. Gonzalez-Hernandez, J. Phys. Chem. Solids, 1999, 60, 1729 CrossRef CAS.
- A. Mendoza-Galvan, J. F. Perez-Robles, F. J. Espinoza-Beltran, R. Ramirez-Bon, Y. V. Vorobiev and J. Gonzalez-Hernandez, J.
Vac. Sci. Technol. A, 1999, 17, 1103 CrossRef CAS.
- E. M. B. de Sousa, A. O. Porto, P. J. Schilling, M. C. M. Alves and N. D. S. Mohallem, J. Phys. Chem. Solids, 2000, 61, 853 CrossRef CAS.
- G. De, J.
Sol–Gel Sci. Technol., 1998, 11, 289 Search PubMed.
- G. De, M. Gusso, L. Tapfer, F. Gonella, G. Mattei, P. Mazzoldi and G. Battaglin, J. Appl. Phys., 1996, 80, 6734 CrossRef CAS.
- C. J. Brinker and G. W. Scherer, Sol–gel
science: the physics and chemistry of sol–gel processing, Academic
Press, New York, 1990. Search PubMed.
- T. Rajh, M. I. Vucemilovic, N. M. Dimitrijevic, O. I. Micic and A. J. Nozik, Chem. Phys. Lett., 1988, 143, 305 CrossRef CAS.
- M. Nogami, S. Suzuki and K. Nagasaka, J. Non-Cryst.
Solids, 1991, 135, 182 Search PubMed.
- L. Spanhel, E. Arpac and H. Schmidt, J. Non-Cryst. Solids, 1992, 147(148), 657 Search PubMed.
- P. Lefebvre, T. Richard, J. Allegre, H. Mathieu, A. Pradel, J. L. Marc, L. Boudes, W. Granier and M. Ribes, Superlattices Microstruct., 1994, 15, 447 CrossRef CAS.
- O. Levy and L. Esquivias, Adv. Mater., 1995, 7, 120.
- V. S. Gurin, V. B. Prokopenko, I. M. Melnichenko, E. N. Poddenezhny, A. A. Alexeenko and K.
V. Yumashev, J. Non-Cryst. Solids, 1998, 232–234, 162 Search PubMed.
- M. Ristov, G. Sinadinovski and I. Grozdanov, Thin Solid Films, 1985, 123, 63 CrossRef CAS.
- R. J. Araujo, J. Butty and N. Peyghambarian, Appl. Phys. Lett., 1996, 68, 584 CrossRef CAS.
- J. Ghijsen, L. H. Tjeng, J. van Elp, H. Eskes and M. T. Czyzyk, Phys. Rev. B, 1988, 38, 11322 CrossRef CAS.
- F. Marabelli, G. B. Parravicini and F. Salghetti-Drioli, Phys. Rev. B, 1995, 52, 1433 CrossRef CAS.
- U. Kreibig
and M. Vollmer, Optical
Properties of Metal Clusters, Springer, Berlin, 1995. Search PubMed.
- R. H. Magruder III, R. F. Haglund
Jr., L. Yang, J. E. Wittig and R. A. Zuhr, J. Appl. Phys., 1994, 76, 708 CrossRef.
- P. Mulvaney, Langmuir, 1996, 12, 788 CrossRef CAS.
- K. Fukumi, A. Chayahara, K. Kadono, H. Kageyama, N. Kitamura, H. Mizoguchi, Y. Horino and M. Makihara, J. Surf. Anal., 1998, 4, 214 Search PubMed.
- N. I. Vlasova, L. I. Demkina and V. I. Karaseva, Fiz.
Khim. Stekla (St. Petersburg), 1984, 10, 345 Search PubMed.
- C. L. Ballhausen, Introduction to ligand field theory, McGraw-Hill
Book Company, Inc., New York, 1964. Search PubMed.
- M. J. Mulder, Phys. Status Solidi (A), 1973, 15, 409 Search PubMed.
- R. Marshall and S. S. Mitra, J. Appl. Phys., 1965, 36, 3882 CAS.
- P. S. McLeod, L. D. Patrain, D. E. Sawyer and T.
M. Peterson, Appl. Phys. Lett., 1984, 45, 472 CrossRef CAS.
- N.
A. Vlasenko and Ya.F. Kononets, Ukr. Fiz. Zh., 1971, 16, 237 Search PubMed.
- K. L. Chopra and S. R. Das, Thin
Film Solar Cells, Plenum Press, New York, 1983. Search PubMed.
- G. P. Sorokin, Yu. M. Papshev and P. T. Oush, Fiz.
Tverd. Tela (St. Petersburg), 1965, 7, 2245 Search PubMed.
- V. I. Klimov, P. H. Bolivar, H. Kurz, V. Karavanskii, V. Krasovskii and Yu. Korkishko, Appl. Phys. Lett., 1995, 67, 653 CrossRef CAS.
- V. I. Klimov, P. H. Bolivar, H. Kurz and V. Karavanskii, Superlattices Microstruct., 1996, 20, 399 CrossRef CAS.
- V. I. Klimov and V. A. Karavanskii, Phys.
Rev. B, 1996, 54, 8087 CrossRef CAS.
- A. D. Yoffe, Adv. Phys., 1993, 42, 173 CAS.
- S. V. Gaponenko, Optical
properties of semiconductor nanocrystals, Cambridge University
Press, Cambridge, 1998. Search PubMed.
- V. S. Gurin, V. B. Prokopenko, A. A. Alexeenko, D. L. Kovalenko and I. M. Melnichenko, J. Inclus.
Phenom. Macrocycl. Chem., 1999, 35, 291 Search PubMed.
- V. S. Gurin, V.
B. Prokopenko, A. A. Alexeenko, I. M. Melnichenko, V. P. Mikhaolov, K. V. Yumashev and A. M. Malyarevich, Funct.
Mater. (Kharkov), 1999, 6, 464 Search PubMed.
- Z. Vucic, O. Milat, V. Horvatic and Z. Ogorelec, Phys. Rev. B, 1981, 24, 5398 CrossRef CAS.
-
(a) R. M. Murray and R. D. Heyding, Can. J. Chem., 1975, 53, 878 CAS;
(b) R. M. Murray and R. D. Heyding, Can. J. Chem., 1976, 54, 841.
- U. Woggon, Optical Properties of Semiconductor
Quantum Dots, Springer-Verlag, Berlin,
Heidelberg, New York, 1997. Search PubMed.
Footnote |
† Basis of a presentation given
at Materials Discussion No. 3, 26–29 September, 2000, University
of Cambridge, UK. |
|
This journal is © The Royal Society of Chemistry 2001 |
Click here to see how this site uses Cookies. View our privacy policy here.